Pd Nanoparticles and Aminopolymers Confined in Hollow Silica

Jan 22, 2019 - Site-directed Immobilization of Bilirubin Oxidase for Electrocatalytic Oxygen Reduction. ACS Catalysis. Al-Lolage, Bartlett, Gounel, St...
0 downloads 0 Views 1MB Size
Subscriber access provided by Gothenburg University Library

Article

Pd Nanoparticles and Aminopolymers Confined in Hollow Silica Spheres as Efficient and Reusable Heterogeneous Catalysts for Semihydrogenation of Alkynes Yasutaka Kuwahara, Hiroto Kango, and Hiromi Yamashita ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.8b04653 • Publication Date (Web): 22 Jan 2019 Downloaded from http://pubs.acs.org on January 26, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 15 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Pd Nanoparticles and Aminopolymers Confined in Hollow Silica Spheres as Efficient and Reusable Heterogeneous Catalysts for Semihydrogenation of Alkynes Yasutaka Kuwahara,†,‡ Hiroto Kango,† and Hiromi Yamashita*,†,‡ Division of Materials and Manufacturing Science, Graduate School of Engineering, Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, Japan †

Unit of Elements Strategy Initiative for Catalysts & Batteries (ESICB), Kyoto University, Katsura, Kyoto 615-8520, Japan ‡

ABSTRACT: A yolk-shell nanostructured composite composed of Pd nanoparticles (NPs) and aminopolymers, poly(ethyleneimine) (PEI), confined in hollow silica spheres which act as an efficient and stable heterogeneous catalyst for semihydrogenation of alkynes is reported herein. The yolk–shell nanostructured Pd-PEI-silica composite catalysts (Pd+PEI@HSS), consisting of Pd NPs core ca. 5−9 nm in diameter and a porous silica shell ca. 30−50 nm in shell thickness, are fabricated by a facile one-pot method using linear- or branched-type PEI (Mw = 1,800~2,500) as an organic template. On the basis of comprehensive structural analyses by FE-SEM, TEM, N2 physisorption, IR, TG, and Pd K-edge XAFS, we show that metal Pd NPs and PEI molecules are encapsulated in the hollow silica sphere with a size of ca. 100−160 nm. The Pd+PEI@HSS composite shows an activity at near room temperature in the liquid-phase hydrogenation of diphenylacetylene to selectively produce cis-stilbene with 95% yield, which outperforms those of the previously-reported Pd/PEI and Lindlar catalysts. Interestingly, the catalyst encapsulating linear-type PEI provides a markedly high alkene selectivity in the semihydrogenation of phenylacetylene to produce styrene owing to the strong poisoning effect of linear PEI, which is clearly revealed by an isotope study using H2/D2/acetylene (or ethylene) gases. The catalyst synthesized with optimum silica shell thickness can be easily recovered and recycled without ant loss of palladium species and PEI and retaining high activities and selectivities over multiple cycles owing to the ability of the protective effect of silica shell, rendering this material an efficient and stable catalyst for semihydrogenation of alkynes.

KEYWORDS: yolk-shell nanostructure, porous materials, Pd nanoparticles, aminopolymer, semihydrogenation

1. INTRODUCTION Selective hydrogenation of alkynes to produce alkenes is an important and fundamental reaction for the synthesis of commodity chemicals and fine chemicals.1-3 Although a large number of homogeneous/heterogeneous catalysts have been developed for semihydrogenation of alkynes,4-5 overwhelming majority employed for liquid-phase semihydrogenation of alkynes is Pd nanoparticle (NP) catalyst.6-8 The Lindlar catalyst (Pd/CaCO3 treated by Pb salts) has long been utilized as a benchmark heterogeneous catalyst.9-10 Lead is used as a modifier to promote selective reduction of alkynes to cis-alkenes, and a basic compound such as quinoline is often added as an organic modifier in the reaction solution to retard the subsequent hydrogenation of cis-alkenes to alkanes; however, this catalyst suffers from some critical drawbacks such as high toxicity of Pb and the low alkene selectivity toward terminal alkynes. Development of alternative, greener catalysts while achieving high alkene selectivity and acceptable reaction rate has been a grand challenge for this reaction.

Owing to the strong hydrogenation ability of Pd, alkynes are readily over-hydrogenated to give the corresponding undesirable alkanes over naked Pd NPs.11 Selectivity can be improved by controlling the adsorption kinetics of the reactants (i.e. alkynes and alkenes) on Pd surface, which can typically be manipulated by introducing organic modifiers containing N or S atoms that can strongly coordinate on Pd NPs.12-15 For example, Pd NPs stabilized with alkylthiol surfactants show different selectivity in the liquid-phase hydrogenation of alkynes and allyl alcohols.16-17 Addition of dimethyl sulfoxide (DMSO) is known to improve the alkene selectivity by acting as homogeneous modifiers.18-19 Amine-based dendrimers have been used as effective scaffolds to immobilize Pd NPs to demonstrate selective hydrogenation of alkynes.20-21 Recently, poly(ethyleneimine) (PEI) emerged as promising supports for Pd NP catalysts.2224 Abundant amine groups in its polymer chain efficiently coordinate and immobilize Pd NPs, enabling chemoselective semihydrogenation of both internal and terminal alkynes to afford the corresponding alkenes in the presence of 1 atm H2.22-23 Although these surfactant-capped and polymer-stabilized Pd NPs show efficient and tunable cata-

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

lytic properties due to their adequate coordination efficiency, the inherent difficulties in separation and leaching of modifiers as well as sintering of Pd species during catalytic use limit their practical application. Recently, much effort has been directed toward development of recoverable and reusable heterogeneous catalytic systems with good activity and selectivity.3 Silica has typically been utilized as an inert oxide support to immobilize Pd NPs with good dispersion and stability.21, 25-26 Mitsudome et al. synthesized silica-supported Pd NPs covered with alkyl sulfoxide groups-containing silica shell for selective semihydrogenation of internal/terminal alkynes.27 Amine-functionalized silicas have been used to immobilize Pd NPs due to a strong metal-support interaction; however, high alkene selectivity was unachievable without the addition of homogeneous organic modifiers.19, 28 Jones et al. synthesized Pd-PEI-silica composite catalyst by immobilizing Pd NPs into a SBA-15 mesoporous silica functionalized with PEI polymer, which showed excellent activity and selectivity in the liquid- phase hydrogenation of diphenylacetylene. 29 However, leaching of amine-functional groups and Pd species during catalytic use through open-ended pores and the associated catalyst deactivation still remain as a major concern. Herein, we envisioned that hollow silica spheres having enclosed cavity spaces could be a reasonable option to immobilize Pd NPs together with PEI, since the silica shell is expected to act as a protective barrier to circumvent the sintering of Pd NPs and leaching of PEI during the catalysis, leading to an improved stability and reusability of catalyst. Hollow silica encapsulating catalytic components in their internal void spaces (so-called “yolk–shell” nanostructures) is emerging category of nanostructured materials.30-37 In most studies on yolk-shell nanostructured catalysts, the enclosed inner void spaces are used for encapsulation of active catalytic components (such as metal NPs,38-43 metal oxides,44-46 and metal complexes47-48); however, encapsulation of bulky functional polymers for improvement of catalytic performances has rarely been examined. Although a number of strategies have been developed for the fabrication of yolk-shell silica nanostructures, most of them requires complex multiple synthetic procedures, such as ship-in-bottle,49 hard templating,50-51 and selective etching approaches.38, 44, 52-53 Furthermore, selective introduction of PEI polymer with large molecular weight within the confined nanospace of hollow silica via a post-synthetic approach is quite difficult.39, 54 In this context, He et al. recently developed a facile one-pot method to encapsulate metal NPs in hollow silica spheres with a size of about 100 nm using poly(acrylic acid) (PAA) as a sole organic template.55 PAA acted as a core template and as a stabilizing agent of the metal NPs through the coordination interaction between the carboxylate groups on the chain ends and the empty orbital of the metals, and a silica shell was subsequently created around the PAA-metal NP aggregate to form yolk-shell nanostructured silica particle. This precedent work inspired us to develop a new method to synthesize hollow silica spheres encapsulating Pd NPs and PEI using PEI as an organic template, since PEI also has an ability

Page 2 of 15

to strongly coordinate on metal NPs due to the presence of lone pairs on N atoms of amines.22-23 In this study, we report a synthesis of new type of yolkshell nanostructured composite composed of Pd NPs and aminopolymers, PEI, confined in hollow silica spheres (Pd+PEI@HSS) for selective hydrogenation of alkynes to produce alkenes. The yolk–shell nanostructured catalyst was for the first time fabricated via a facile one-pot synthesis by using linear or branched PEI as a self-template. The resulting composite was found to act as an efficient and reusable catalyst in the semihydrogenation of internal/terminal alkynes in the presence of 1 atm of H2 while retaining high alkene selectivity and minimizing the unfavorable over-hydrogenation, which far outperformed those achieved on the prototype supported Pd NP catalysts. Kinetic analysis and isotope experiment were examined to understand the role of PEI in tuning the alkene selectivity. Furthermore, leaching and reusability tests were performed to verify the ability of the silica shell to protect the encapsulated components.

2. EXPERIMENTAL SECTION 2.1. Chemicals and Materials. Tetraethoxy orthosilicate (TEOS, 95%), disodium tetrachloropalladate(II) (Na2PdCl4, >95%), palladium(II) chloride (PdCl2, 99%), palladium(II) acetate (Pd(OAc)2, >97%), sodium borohydride (NaBH4, >95%), aqueous ammonia solution (28%), methanol (99.8%), ethanol (99.5%), 1,4-dioxane (99.5%) and 8-hexadecyne (>99%) were purchased from Nacalai Tesque Inc. Poly(ethyleneimine) (linear, Mw = 2,500, 99%) and diphenylacetylene (98%) was purchased from Sigma-Aldrich. Poly(ethyleneimine) (branched, Mw = 600, 1,800, and 10,000, 99%) was purchased from Alfa Aesar. Phenylacetylene (>97%), methyl phenylpropiolate (>98%), ethyl phenylpropiolate (>97%), phenylpropiolic acid (>98%) and Lindlar catalyst (5% Pd on CaCO3 poisoned with 2.4-3.0% Pb) as a reference catalyst were purchased from Tokyo Chemical Industry Co., Ltd. 20% H2/Ar, 20% D2/Ar, 10% C2H2/Ar, and 10% C2H4/Ar gas used for gas-phase hydrogenation experiment were purchased from Taiyo Nippon Sanso Corp. 2.2. Characterization. Field-emission scanning electron microscopy (FE-SEM) images were taken with a JEOL JSM6500F operating at 12.0 kV. Transmission electron microscopy (TEM) images were taken with a Hitachi H-800 FETEM operating at 200 kV. Scanning transmission electron microscope (STEM) images and elemental maps were collected on a FEI TITAN80-300 instrument operated with an accelerating voltage of 200 kV and quipped with an EDAX r-TEM/SuperUTW energy-dispersive X-ray detector. The Pd loadings of the samples were determined by inductively coupled plasma atomic emission spectrometry (ICP-AES). The PEI content of the samples were determined from the thermogravimetric (TG) analysis data measured on a Rigaku Thermo plus EVO2 under a flow of air (30 mL/min) with a ramping rate of 10 °C/min. Infrared (IR) spectra were collected on a JASCO FT/IR-6300 instrument under vacuum using samples diluted with KBr. Nitrogen adsorption–

ACS Paragon Plus Environment

Page 3 of 15 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

desorption isotherms were measured at –196 °C with a MicrotracBEL BELSORP-max system. Prior to measurements, samples were degassed at 120 °C under vacuum for 3 h to vaporize the physisorbed water. The surface area was calculated from the adsorption data in the range p/p0 = 0.050.25 using the Brunauer–Emmett–Teller (BET) method, and the pore size distributions were calculated by using the Barrett–Joyner–Halenda (BJH) method. Powder X-ray diffraction (XRD) patterns were recorded on a Rigaku Ultima IV diffractometer with CuKα radiation (λ = 1.54056 Å). The X-ray absorption fine structure (XAFS) measurements for Pd K-edge were performed in fluorescence mode at the BL01B1 beamline at the SPring-8 (JASRI), Hyogo, Japan. A double Si(111) crystal monochromator was used for energy selection and a Pd foil was used for energy calibration. Fluorescence Pd K-edge XAFS spectra were collected at room temperature with a Lytle detector. The extended X-ray absorption fine structure (EXAFS) data were normalized by fitting the background absorption coefficient. Fourier transformation of k3-weighted normalized EXAFS data (FT-EXAFS) from k space to r space was performed over the range of 3.0 ≤ k (Å−1) ≤ 11.0 to obtain the radial structure functions (RDFs). 2.3. Preparation of Pd-PEI-Silica Composites (Pd+PEI@HSS). Yolk-shell nanostructured Pd+PEI@HSS composites were synthesized by an one-pot method developed based on Stöber method55-56 using linear- or branched PEI as an organic template, as schematically shown in Scheme 1. In a typical procedure, 0.12 g of PEI was dissolved in aqueous ammonia (2.8 mL, 28%) with magnetic stirring at room temperature, and 0.8 mL of aqueous solution of Na2PdCl4 (20 mg/mL) and 90 mL of EtOH were subsequently added, followed by vigorous stirring for 30 min to obtain a transparent solution (Scheme 1, step 1; see also the Supporting Information, Figure S1a). Next, 2.0 mL of freshly-prepared aqueous NaBH4 solution (20 mg/mL) (approximately 20 equiv. of Pd) was slowly added to obtain brownish suspension (step 2; see also Figure S1b). To this solution was rapidly added 1.8 mL of TEOS as a Si source under vigorous stirring (step 3), followed by continuous stirring for 6 h at room temperature to ensure silica shell formation (step 4; see also Figure S1c). The resulting precipitate was recovered by centrifugation, washed with DI water (50 mL) and EtOH (50 mL), and dried under vacuum overnight to afford Pd+PEI@HSS as a greyish powder. To investigate the impacts of PEI type on the structure and the catalytic performance of the final solid, synthesis was performed by using linear PEI (Mw = 2,500) and branched PEI (Mw = 1,800), which were denoted as Pd+PEI(L)@HSS and Pd+PEI(B)@HSS, respectively. Furthermore, these samples were calcined at 550 °C in air for 5 h to remove organic components, which were then reduced at 200 °C in a flow of H2 (20 mL/min) to give Pd(0) NPs encapsulated in hollow silica spheres (denoted as Pd+PEI(L/B)@HSS_cal). To obtain Pd+PEI@HSS with different silica shell thickness, synthesis was also performed

by adding varied amounts of TEOS (0.9, 1.35, 1.8, and 2,7 mL) using linear PEI as a template polymer. The thus obtained samples were denoted as Pd+PEI(L)@HSS_X, where X represents the volume of TEOS added in mL. 2.4. Preparation of Pd/PEI. PEI-supported Pd(0) NPs were synthesized according to the method previously reported in literature.22 Typically, 0.12 g of linear PEI (Mw = 2,500) was completely dissolved in EtOH (90 mL) with magnetic stirring at room temperature, and 0.8 mL of aqueous solution of Na2PdCl4 (20 mg/mL) was added in Ar atmosphere, followed by ultrasonication and stirring for 10 min. The resulting solution was stirred in a continuous flow of H2 (10 mL/min) at room temperature for 24 h to yield a black colloidal solution. After removal of EtOH by evaporation, a black gummy Pd/PEI was obtained. The Pd loading is 4.6 wt% based on the incipient ratio of Na2PdCl4 and PEI. 2.5. Liquid-phase Semihydrogenation of Alkynes. Catalytic reactions were performed at 30 °C in a schlenk flask connected to a water-cooling condenser under a continuous flow of 1 atm H2. In a typical reaction procedure, as-prepared Pd+PEI@HSS (ca. 0.05 g: Pd 0.005 mmol) was placed in the flask, and the reactor was purged with H2 gas for 10 min. Next, alkyne (1.0 mmol) dissolved in MeOH (5 mL) and 1,4-dioxane (5 mL) was injected into the flask via a syringe to initiate the reaction. During the reaction, the reaction mixture was magnetically stirred (600 rpm) at 30± 0.5 °C under a continuous flow of 1 atm H2 (10 mL/min). A small aliquot of reaction solution was withdrawn and the amounts of reactant and products were quantified by a gas chromatograph (GC; Shimadzu, GC-14B) equipped with a flame ionization detector (FID) and a Zebron ZB-WAX plus capillary column using biphenyl as an internal standard. To examine the catalyst reusability, the spent catalyst was recovered by centrifugation, washed with MeOH, dried under vacuum at room temperature, and then subjected to multiple catalytic cycles. 2.6. Gas-phase Hydrogenation Experiment. Gasphase hydrogenation of acetylene and ethylene in the flow of H2/D2 gases was carried out at 120 °C in a temperatureprogrammed desorption (TPD) system (BELCAT-II, MicrotracBEL Corp.). About 0.05 g of catalyst (ca. Pd 0.005 mmol) was mounted in a pyrex reactor and pre-reduced at 12o °C under H2 flow (20 mL/min) for 30 min. In the initial step, mixed gas of H2 and D2 balanced with Ar (H2:D2 = 1:1 volume ratio, 4.3 kPa H2, 4.3 kPa D2, 91.4 kPa Ar) was flowed (total flow: 50 mL/min) under the same temperature condition to check H2-D2 isotope exchange reaction. After approximately 10 min reaction, either C2H2 or C2H4 gas was flowed simultaneously with H2/D2/Ar ((H2+D2)/C2H2 (or C2H4) = 1.5 volume ratio, 4.3 kPa H2, 4.3 kPa D2, 2.9 kPa C2H2 (or C2H4), 88.5 kPa Ar) (total flow: 50 mL/min). The generated species (m/z = 3 for HD, m/z = 2830 for ethylene isotopes, m/z = 30-34 for ethane isotopes) were monitored using an online mass spectrometer (MicrotracBEL BELMass)

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Step 1

Step 2 Pd NPs + NaBH4

Pd2+ Pd2+

Pd NPs

porous SiO2 PEI

+ TEOS

Aging

Pd2+

Pd ions + PEI

Pd+PEI(L/B)@HSS

Pd NPs-PEI aggregates

Template H2N

Step 4

Step 3

PEI Pd2+

Page 4 of 15

N H

H N n

N

NH2

N H

N

Et

NH2 N

NH2

Poly(ethyleneimine) (Linear)

Si source

H N

NH2

N H

Et n

Poly(ethyleneimine) (Branched)

O

O Si O O

Et

=

Et

Tetraethoxy orthosilicate (TEOS)

Scheme 1. Schematic representation of the synthetic procedure of hollow silica spheres encapsulating Pd NPs and poly(ethyleneimine) (Pd+PEI@HSS).

3. RESULTS AND DISCUSSION 3.1. Preparation and Pd+PEI@HSS composites.

Characterization

of

Scheme 1 schematically depicts the synthetic procedure of Pd+PEI@HSS composites using linear- or branched PEI as an organic template. In the first step, a transparent solution was obtained after mixing of PEI and Pd precursor in aqueous ammonia solution (Figure S1a), which immediately turned brown after the addition of NaBH4 (Figure S1b), indicating that PdCl42- ions entrapped by the nitrogen ligands on the polymer chain are reduced to form Pd NPPEI aggregates. In a following step, a brownish milky suspension was obtained in a few minutes after the addition of TEOS as a Si precursor (Figure S1c), representing the condensation of TEOS to form silicate network around the Pd-PEI aggregates as nuclei, which is catalyzed by NH3 added in the solution as a base. The supernatant solution collected after the recovery of the solid was colorless, suggesting that nearly all of Pd had been incorporated in the product. (a)

(b)

500 nm

(d)

(c)

100 nm

(e)

500 nm

100 nm

(f)

100 nm

100 nm

Figure 1. (a, d) FE-SEM images, (b, e) TEM images and (c, f) magnified TEM images of Pd+PEI@HSS synthesized with (a– c) linear PEI (Mw = 2,500) and (d–f) branched PEI (Mw = 1,800).

FE-SEM images show that Pd+PEI@HSS is composed of spherical silicas with an average particle size of ca. 155 nm

and 107 nm for Pd+PEI(L)@HSS and Pd+PEI(B)@HSS, respectively, which are interconnected with each other (Figure 1a and d). TEM images clearly show that the silica spheres have hollow nanostructure with an silica shell thickness of ca. 49 nm and 36 nm for Pd+PEI(L)@HSS and Pd+PEI(B)@HSS, respectively, and Pd NPs are observed in their hollow spaces (Figure 1b,c,e and f) (for the distribution diagrams of silica particle size, silica shell thickness, and Pd NP size, see Figure S2). Hollow silica particles without encapsulating Pd NPs were scarcely observed. This indicates that silica formation is directed by the preformed Pd-PEI aggregates as nuclei as well as the electrostatic interaction between amine groups on the polymer chain and TEOS.55 Branched-type PEI afforded smaller silica particles with thinner silica shells, which might be due to the lower molecular weight of PEI than that of linear PEI used for synthesis. The average diameter of Pd NPs was determined to be ca. 8.9 nm and 6.0 nm for Pd+PEI(L)@HSS and Pd+PEI(B)@HSS, respectively; however, no clear characteristic peaks for Pd species were observed in XRD patterns, indicating that Pd NPs are highly dispersed (for XRD patterns, see Figure S3). The morphology of final solid was strongly dependent on the molecular weight of PEI and the kind of Pd precursor employed; branched PEI with smaller molecular weight (Mw = 600) gave solid silica spheres without hollow cavities (Figure S4a). Use of branched PEI with larger molecular weight (Mw = 10,000) resulted in the formation of amorphous silica without any defined nanostructure (Figure S4c). When other Pd(II) precursors, such as H2PdCl4 or Pd(OAc)2 were used, aggregates of small silica particles with ununiformlly-distributed Pd NPs were obtained (see Figure S5). Thus, appropriate choice of PEI and Pd precursor is needed to synthesize yolk-shell nanostructured Pd-PEI-silica composite, since the coordination interaction between Pd species and amine sites on PEI polymer differs depending on the kind of Pd precursor, and the size of Pd-PEI aggregate can be varied depending on the molecular weight of PEI.

ACS Paragon Plus Environment

Page 5 of 15 (a)

(b)

(c)

Pd

(f)

N

direction

(d)

Si (e)

O

(g)

Pd NPs

PEI

Si O C N Pd

Intensity / a.u.

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

0

20 40 60 80 100 120 140 160

Distance / nm

Figure 2. (a) STEM image, (b) High-angle annular dark-field STEM (HAADF-STEM) image, (c-f) the corresponding STEM elemental maps of (c) Pd, (d) Si, (e) O, and (f) N of Pd+PEI(L)@HSS. (g) STEM elemental line scan across the Pd+PEI(L)@HSS particle in (b).

Table 1. Compositional and Textural Properties of Pd+PEI@HSS Composites Samples Pd+PEI(L)@HSS Pd+PEI(L)@HSS_cal Pd+PEI(B)@HSS

PEI type

Pd contenta (wt%)

PEI contentb (wt%)

Linear, Mw = 2,500

1.1

24.0

-

1.5

Branched, Mw = 1,800

1.1

N2 physisorption SBET

Vtotald(cm3/g)

De (nm)

34

0.12

2.7

0

19

0.09

2.8

21.9

35

0.16

2.8

c

(m2/g)

Pd+PEI(B)@HSS_cal

-

1.4

0

28

0.15

2.6

Lindlar catalystf

-

5.0

0

30

0.09

2.7

a Determined by ICP-AES. b Determined by TG analysis. c Surface area calculated by BET method. d Total pore volume reported at p/p0=0.99. e Peak pore size determined by BJH method. f 5% Pd on CaCO3 (poisoned with 3% Pb).

In the HAADF-STEM micrograph of Pd+PEI(L)@HSS (Figure 2), Pd NPs with more electron density located inside the hollow cavities, mostly on the inside silica wall, are clearly observed (Figure 2b). Whole area mapping analysis visualizes that the distributions of Si atoms with blue dots (Figure 2d) and O atoms with pink dots (Figure 2e) overlap with the shell region, and Pd atoms with yellow dots (Figure 2c) are seen in the hollow region, again evidencing the formation of yolk-shell nanostructure consisting of Pd NPs as a core and silica as a shell. N atoms with white dots (Figure 2f) are uniformly distributed throughout the particles, indicating that the PEI is incorporated in both shell and hollow regions. The line mapping of a selected silica particle confirms the inclusion of N, C and Pd atoms in the same axial region (Figure 2g), suggesting a close proximity betwen Pd NPs and PEI. In FT-IR spectra of Pd+PEI@HSS (Figure 3A), the IR bands at 2962 and 2858 cm−1 assigned to the stretching vibration of aliphatic C−H bond, the band at 1467 cm−1 assigned to the bending vibration of C−H bond, and the peak at 1560 cm−1 assignable to the bending vibration of N−H bond are observed, which completely disappear after thermal treatment at 550 °C in air, further ascertaining the successful incorporation of PEI in the as-synthesized material. The PEI contents were estimated to be 24.0 and

21.9 wt% for Pd+PEI(L)@HSS and Pd+PEI(B)@HSS, respectively, from thermogravimetric (TG) data (Figure 3B), and the Pd contents were determined to be 1.1 wt% in both cases (see Table 1) by elemental analysis, confirming the majority of Pd and PEI added during the synthsis was incorporated in the final solid. The porous structure of Pd+PEI@HSS materials was investigated by N2 physisorption measurement. A clear hysteresis observed in the range of 0.4 < p/p0 < 1.0 in N2 adsorption-desorption isotherms indicates the existence of hollow cavities surrounded by mesopore systems (Figure 3C).57 The pore size distributions calculated by the BJH method confirm the existence of broadly-distributed mesopores with average pore size of ca. 2.7 nm (Figure 3D) probably due to the formation of hydroxylated amorphous silica network. The textural properties obtained from the N2 physisorption analysis are listed in Table 1. BET surface area (SBET) and total pore volume (Vtotal) were determined to be 34 m2/g and 0.12 cm3/g for Pd+PEI(L)@HSS and 35 m2/g and 0.16 cm3/g for Pd+PEI(L)@HSS, showing that the type of PEI had little influence on the porous structures of the final solid. Nitrogen adsorption data of calcined samples (i.e. Pd+PEI(L)@HSS_cal and Pd+PEI(L)@HSS_cal) shows substantial decreases in surface area and pore volume (Table 1), which might be due

ACS Paragon Plus Environment

ACS Catalysis

5

Pd+PEI(L)@HSS Pd+PEI(B)@HSS

0

(a)

Weight loss / wt%

Transmittance / a.u.

(B)

(b)

(c)

-5 -10 -15

3000

2000

-30 0

1000

24300

(C) 150

dVp/ddp / cm3・g-1・nm-1

100 +50

50

0 0.0

0.2

0.4

0.6

p/p0

0.8

1.0

2.7 nm

Pd+PEI(L)@HSS Pd+PEI(B)@HSS

2.8 nm

+0.01

0.020 0.015

24400

24450

Pd+PEI(L)@HSS Pd+PEI(B)@HSS PdO Pd(NH3)2Cl2

0

1

2

3

4

5

6

Interatomic distance / Å

0.005

5

10

15

20

25

Figure 4. (A) Pd K-edge XANES spectra and (B) radial distribution functions (RDFs) obtained from Pd K-edge EXAFS oscillations for Pd+PEI@HSS composites and reference Pd samples.

3.2. Catalytic Performance of Pd+PEI@HSS composites.

0.010

0.000 0

24350

Pd foil Pd-N Pd-O

100 200 300 400 500 600

(D) 0.025

Pd+PEI(L)@HSS Pd+PEI(B)@HSS

Pd foil Pd+PEI(L)@HSS Pd+PEI(B)@HSS PdO Pd(NH3)2Cl2

Photon energy / eV Temperature / ℃

Wavenumber / cm-1

Pd-Pd

-20 -25

4000

(B) FT magnitude / a.u.

(A)

(A) Normalized absorption / a.u

to the thermal shrinkage of silica network to create denser silica shells (for the comparison of N2 physisorption isotherms of as-synthesized and calcined samples, see Figure S6).

Volume adsorbed / cm3 (STP) g-1

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 15

30

Pore diameter / nm

Figure 3. (A) FT-IR spectra of (a) PEI(L), (b) Pd+PEI(L)@HSS, and (c) Pd+PEI(L)@HSS_cal. (B) Weight loss curves, (C) N2 adsorption-desorption isotherms, and (D) the corresponding pore size distribution curves determined by BJH method for as-synthesized Pd+PEI@HSS composites.

Figure 4A and 4B show X-ray absorption near-edge spectra (XANES) and RDFs obtained from the k3-weighted Pd K-edge EXAFS oscillations of the prepared samples, respectively, together with those of reference Pd samples. The X-ray absorption edges of the Pd+PEI@HSS samples (E0 = 24355.5 eV) are positioned between the absorption edges of Pd(0) foil (E0 = 24352.1 eV) and Pd(II)O (E0 = 24358.3 eV), and the height of white line intensity increases in the order of Pd(0) foil < Pd+PEI(L/B)@HSS < Pd(II)O (Figure 4A), indicating the coexistence of Pd(0) and Pd(II) species in the Pd+PEI@HSS samples. In the RDFs of Pd Kedge EXAFS spectra, both Pd+PEI@HSS samples exhibit two distinct peaks at around r = 1.6 and 2.5 Å (the phase shift uncorrected); the first shell at around r = 1.6 Å is associated with the backscattering due to the neighboring nitrogen or oxygen atoms and the second shell at around r = 2.5 Å is attributable to the Pd–Pd bond (Figure 4B). Since the peak ascribed to Pd–O–Pd bond, which typically appear at around r = 3.0 Å, is scarcely observed, the former peak (at r = 1.6 Å) can dominantly be ascribed to the Pd(II) species ligated to the nitrogen moieties of the polymer framework.58 These results, combined with elemental mapping analysis, indicate that the Pd species encapsulated inside hollow silica are primarily present as Pd(0) NPs of which surface is closely surrounded by PEI polymers. The XAFS result also suggests that the the oxidation state of Pd species does not markedly vary depending on the type of PEI used.

The synthesized Pd+PEI@HSS catalysts were initially assessed in the semihydrogenation of diphenylacetylene (1) with 0.5 mol% Pd in a flow of atmospheric pressure of H2 at 30 °C. The reaction profiles over the Pd+PEI@HSS catalysts, the calcined analogues and some supported Pd catalysts are shown in Figure 5. The hydrogenation of diphenylacetylene readily took place without any induction period, indicating that Pd species encapsulated in Pd+PEI@HSS samles are predominantly present as metallic Pd(o) NPs as confirmed by Pd K-edge XAFS. Surprisingly, the as-synthesized Pd+PEI(L/B)@HSS catalysts both selectively afforded cis-stilbene (2) with >90% selectivities due to the syn-addition of H atoms over Pd(o) NPs surface, and trans-stilbene (3) and bibenzyl (4) were hardly produced even after the complete consumption of diphenylacetylene (Figure 5A and C). On the other hand, cis-stilbene was quickly transformed into bibenzyl after the complete conversion of diphenylacetylene over the Pd+PEI(L/B)@HSS calcined in air, PEI-free analogues (Figure 5B and D). This result clearly demonstrates that PEI is capable of efficiently suppressing the over-hydrogenation of stilbene to bibenzyl owing to the strong coordination ability onto Pd NP, which is confirmed by XAFS analysis. It is worth mentioning that the calcined catalysts both showed obviously faster hydrogenation rates compared with the as-synthesized catalysts, due to the removal of PEI covering the surface of Pd NPs. Considering the fact that Pd NPs are the sole active component existing in the present catalytic systems, Pd NPs inside the hollow silica spheres are accessible to reactants through ill-defined mesopores created in the silica shell, and PEI protects the Pd NPs without fully preventing the access to the reactants. Pd+PEI(B)@HSS provided a slightly faster reaction rate compared with Pd+PEI(L)@HSS, which might be explained by the formation of thinner silica shell (36 nm vs. 49 nm) and smaller Pd NP size (6.0 nm vs. 8.9 nm).

Table 2. Comparison of Catalytic Performances of Pd+PEI@HSS Catalysts and Lindlar Catalyst

ACS Paragon Plus Environment

Page 7 of 15

Samples

NPda (mmol/g)

Nsurfb (mmol/g)

0.103

0.0016

Pd+PEI(L)@HSS

Hydrogenation of diphenylacetylene

Hydrogenation of phenylacetylene

R Yc (h-1)

REd (h-1)

RY/RE

R Yc (h-1)

REd (h-1)

RY/RE

4693

99

95

95 : 5

480

>99

81

360

>99

94

300

>99

76

180

>99

100

94 : 6 100 : 0 100 : 0 100 : 0

conditions: catalyst (Pd+PEI(L)@HSS, Pd 0.5 mol%), alkyne (1 mmol), MeOH:1,4-dioxane = 1:1 (10 mL), 30 °C, under a flow of 1 atm H2 (10 mL/min). bDetermined by GC-FID. cEtOH:1,4-dioxane = 1:1 (10 mL) was used as solvent.

after 5th use

1

Time Conv.b Alkene Sel.b (min) (%) (%)

aReaction

Interatomic distance / Å 150

O OEt

0 Volume adsorbed / cm3 (STP) g-1

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

100 200 300 400 500 600

Temperature / ℃

Figure 10. (A) TEM image of Pd+PEI(L)@HSS catalyst after five catalytic runs. (B) Pd K-edge FT-EXAFS spectra, (C) N2 adsorption-desorption isotherms, and (D) weight loss curves

Furthermore, Pd+PEI(L)@HSS_1.8 catalyst, the best performance catalyst among examined in this study, provided good selectivity for the semihydrogenation of other alkynes (Table 4). Several kinds of aromatic and aliphatic internal alkynes were efficiently hydrogenated to the corresponding alkenes with good to excellent yields (entries 2– 6) using atmospheric pressure of H2 as a reductant, while suppressing over-hydrogenation to corresponding alkanes. These results led us to a conclusion that Pd+PEI(L)@HSS catalyst having an optimum silica thickness can act as a stable, reusable, and efficient heterogeneous catalyst in the liquid-phase semihydrogenation of alkynes.

4. CONCLUSIONS

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

A new type of Pd-PEI-silica composite with yolk–shell nanostructure was developed for the selective hydrogenation of alkynes to produce alkenes. The synthesized composite was composed of Pd NPs core less than 9 nm in diameter and linear/branched PEI which were confined in hollow silica spheres having ill-defined mesoporous structures with a particle size of 100-160 nm. The present synthetic protocol provides a facile one-pot synthetic process for the fabrication of yolk-shell silica nanostructure with tunable size and shell thickness without severe control of synthetic conditions. Such a yolk-shell nanostructured composite could act as an efficient and reusable heterogeneous catalyst in the semihydrogenation of alkynes in the presence of 1 atm H2, giving a markedly improved alkene selectivity compared with the previously-reported prototype Pd catalysts. In particular, the catalyst encapsulating linear PEI afforded a high alkene selectivity in the semihydrogenation of phenylacetylene to produce styrene, in which silica shell thickness had a strong influence on alkene selectivity. Comprehensive analysis together with isotope experiment revealed that Pd NPs and PEI polymers are encapsulated within the hollow silica particles in close proximity, in which the PEI acts as a macro-ligand to strongly coordinate on Pd NP surface, enabling discriminative adsorption of alkynes over alkenes. This leads to highly selective hydrogenation of alkynes to alkenes with suppressing over-hydrogenation into alkanes. Owing to the rigidity of the protective silica shell, the catalyst was easily recoverable and reusable without leaching of Pd species and PEI, with retaining high activities and selectivities over multiple cycles, rendering this material an efficient and stable catalyst for semihydrogenation of alkynes.

The synchrotron radiation experiments for XAFS measurements were performed at the BL01B1 beam line in SPring-8 with the approval from JASRI (no. 2018A1089).

REFERENCES 1. 2. 3. 4. 5.

6. 7.

8.

9. 10. 11.

ASSOCIATED CONTENT Supporting Information. This Supporting Information is available free of charge via the Internet at http://pubs.acs.org. Photographs of synthetic solutions (Figure S1), statistical structural data of samples (Figure S2), XRD patterns (Figure S3), TEM images of samples synthesized with different polymers (Figure S4) or different Pd precursors (Figure S5), N2 physisorption data upon calcination (Figure S6), statistical structural data of samples synthesized with varied TEOS amounts (Figure S7), and reaction profiles during catalyst reusability test (Figure S8) (PDF)

12.

13.

14.

AUTHOR INFORMATION 15.

Corresponding Author *[email protected]

Notes The authors declare no competing financial interest.

16.

ACKNOWLEDGMENT This work was supported by a Grant-in-Aid from the Frontier Research Base for Global Young Researchers, Osaka University, and the Grant-in-Aid for Young Scientists (KAKENHI) from JSPS (no. 18K14056). A part of this work was supported by the “Grant for Research” of The japan Petroleum Institute.

Page 12 of 15

17.

Marvell, E. N.; Li, T. Catalytic semihydrogenation of the triple bond. Synthesis 1973, 1973, 457-468. Chinchilla, R.; Najera, C. Chemicals from alkynes with palladium catalysts. Chem. Rev. 2014, 114, 1783-1826. Crespo-Quesada, M.; Cárdenas-Lizana, F.; Dessimoz, A.-L.; Kiwi-Minsker, L. Modern trends in catalyst and process design for alkyne hydrogenations. ACS Catal. 2012, 2, 1773-1786. Laren, M. W. v.; Elsevier, C. J. Selective homogeneous palladium(0)-catalyzed hydrogenation of alkynes to (Z)-alkenes. Angew. Chem. Int. Ed. 1999, 38, 3715-3717. Drost, R. M.; Bouwens, T.; van Leest, N. P.; de Bruin, B.; Elsevier, C. J. Convenient transfer semihydrogenation methodology for alkynes using a PdII-NHC precatalyst. ACS Catal. 2014, 4, 1349-1357. Jackson, S. D.; Shaw, L. A. The liquid-phase hydrogenation of phenyl acetylene and styrene on a palladium/carbon catalyst. Appl. Catal. A 1996, 134, 91-99. Mitsudome, T.; Urayama, T.; Yamazaki, K.; Maehara, Y.; Yamasaki, J.; Gohara, K.; Maeno, Z.; Mizugaki, T.; Jitsukawa, K.; Kaneda, K. Design of core-Pd/shell-ag nanocomposite catalyst for selective semihydrogenation of alkynes. ACS Catal. 2015, 6, 666-670. Lu, Y.; Feng, X.; Takale, B. S.; Yamamoto, Y.; Zhang, W.; Bao, M. Highly selective semihydrogenation of alkynes to alkenes by using an unsupported nanoporous palladium catalyst: No leaching of palladium into the reaction mixture. ACS Catal. 2017, 7, 8296-8303. Lindlar, H. Ein neuer katalysator für selektive hydrierungen. Helv. Chim. Acta 1952, 35, 446–450. Gutman, H.; Lindlar, H. In Chemistry of Acetylenes; Viehe, H. G., Ed.; Marcel Dekker: New York 1969. Chung, J.; Kim, C.; Jeong, H.; Yu, T.; Binh, D. H.; Jang, J.; Lee, J.; Kim, B. M.; Lim, B. Selective semihydrogenation of alkynes on shape-controlled palladium nanocrystals. Chem. Asian J. 2013, 8, 919-925. Yun, S.; Lee, S.; Yook, S.; Patel, H. A.; Yavuz, C. T.; Choi, M. Cross-linked “poisonous” polymer: Thermochemically stable catalyst support for tuning chemoselectivity. ACS Catal. 2016, 6, 2435-2442. Moreno, M.; Ibanez, F. J.; Jasinski, J. B.; Zamborini, F. P. Hydrogen reactivity of palladium nanoparticles coated with mixed monolayers of alkyl thiols and alkyl amines for sensing and catalysis applications. J. Am. Chem. Soc. 2011, 133, 43894397. Crespo-Quesada, M.; Dykeman, R. R.; Laurenczy, G.; Dyson, P. J.; Kiwi-Minsker, L. Supported nitrogen-modified Pd nanoparticles for the selective hydrogenation of 1-hexyne. J. Catal. 2011, 279, 66-74. Kwon, S. G.; Krylova, G.; Sumer, A.; Schwartz, M. M.; Bunel, E. E.; Marshall, C. L.; Chattopadhyay, S.; Lee, B.; Jellinek, J.; Shevchenko, E. V. Capping ligands as selectivity switchers in hydrogenation reactions. Nano Lett. 2012, 12, 5382-5388. Moreno, M.; Kissell, L. N.; Jasinski, J. B.; Zamborini, F. P. Selectivity and reactivity of alkylamine- and alkanethiolate-stabilized Pd and PdAg nanoparticles for hydrogenation and isomerization of allyl alcohol. ACS Catal. 2012, 2, 2602-2613. Gavia, D. J.; Koeppen, J.; Sadeghmoghaddam, E.; Shon, Y. S. Tandem semi-hydrogenation/isomerization of propargyl alcohols to saturated carbonyl analogues by dodecanethiolatecapped palladium nanoparticle catalysts. RSC Adv. 2013, 3, 13642-13645.

ACS Paragon Plus Environment

Page 13 of 15 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

18. Takahashi, Y.; Hashimoto, N.; Hara, T.; Shimazu, S.; Mitsudome, T.; Mizugaki, T.; Jitsukawa, K.; Kaneda, K. Highly efficient Pd/SiO2–dimethyl sulfoxide catalyst system for selective semihydrogenation of alkynes. Chem. Lett. 2011, 40, 405407. 19. Verho, O.; Zheng, H.; Gustafson, K. P. J.; Nagendiran, A.; Zou, X.; Bäckvall, J.-E. Application of Pd nanoparticles supported on mesoporous hollow silica nanospheres for the efficient and selective semihydrogenation of alkynes. ChemCatChem 2016, 8, 773-778. 20. Mizugaki, T.; Murata, M.; Fukubayashi, S.; Mitsudome, T.; Jitsukawa, K.; Kaneda, K. Pamam dendron-stabilised palladium nanoparticles: Effect of generation and peripheral groups on particle size and hydrogenation activity. Chem. Commun. 2008, 241-243. 21. Karakhanov, E.; Maximov, A.; Kardasheva, Y.; Semernina, V.; Zolotukhina, A.; Ivanov, A.; Abbott, G.; Rosenberg, E.; Vinokurov, V. Pd nanoparticles in dendrimers immobilized on silica-polyamine composites as catalysts for selective hydrogenation. ACS Appl. Mater. Interfaces 2014, 6, 8807-8816. 22. Sajiki, H.; Mori, S.; Ohkubo, T.; Ikawa, T.; Kume, A.; Maegawa, T.; Monguchi, Y. Partial hydrogenation of alkynes to cis-olefins by using a novel Pd(0)-polyethyleneimine catalyst. Chem. Eur. J. 2008, 14, 5109-5111. 23. Yabe, Y.; Sawama, Y.; Monguchi, Y.; Sajiki, H. New aspect of chemoselective hydrogenation utilizing heterogeneous palladium catalysts supported by nitrogen- and oxygen-containing macromolecules. Catal. Sci. Technol. 2014, 4, 260-271. 24. Veerakumar, P.; Velayudham, M.; Lu, K.-L.; Rajagopal, S. Silica-supported pei capped nanopalladium as potential catalyst in suzuki, heck and sonogashira coupling reactions. Appl. Catal. A 2013, 455, 247-260. 25. Spee, M. P. R.; Boersma, J.; Meijer, M. D.; Slagt, M. Q.; Koten, G. v.; Geus, J. W. Selective liquid-phase semihydrogenation of functionalized acetylenes and propargylic alcohols with silicasupported bimetallic palladium-copper catalysts. J. Org. Chem. 2001, 66, 1647-1656. 26. Opanasenko, M.; Štěpnička, P.; Čejka, J. Heterogeneous Pd catalysts supported on silica matrices. RSC Adv. 2014, 4, 6513765162. 27. Mitsudome, T.; Takahashi, Y.; Ichikawa, S.; Mizugaki, T.; Jitsukawa, K.; Kaneda, K. Metal-ligand core-shell nanocomposite catalysts for the selective semihydrogenation of alkynes. Angew. Chem. Int. Ed. 2013, 52, 1481-1485. 28. da Silva, F. P.; Fiorio, J. L.; Rossi, L. M. Tuning the catalytic activity and selectivity of Pd nanoparticles using ligand-modified supports and surfaces. ACS Omega 2017, 2, 6014-6022. 29. Long, W.; Brunelli, N. A.; Didas, S. A.; Ping, E. W.; Jones, C. W. Aminopolymer–silica composite-supported Pd catalysts for selective hydrogenation of alkynes. ACS Catal. 2013, 3, 1700-1708. 30. Liu, J.; Qiao, S. Z.; Chen, J. S.; Lou, X. W.; Xing, X.; Lu, G. Q. Yolk/shell nanoparticles: New platforms for nanoreactors, drug delivery and lithium-ion batteries. Chem. Commun. 2011, 47, 12578-12591. 31. Tang, F.; Li, L.; Chen, D. Mesoporous silica nanoparticles: Synthesis, biocompatibility and drug delivery. Adv. Mater. 2012, 24, 1504-34. 32. Wu, S. H.; Mou, C. Y.; Lin, H. P. Synthesis of mesoporous silica nanoparticles. Chem. Soc. Rev. 2013, 42, 3862-3875. 33. Perez-Lorenzo, M.; Vaz, B.; Salgueirino, V.; Correa-Duarte, M. A. Hollow-shelled nanoreactors endowed with high catalytic activity. Chem. Eur. J. 2013, 19, 12196-12211. 34. Li, Y.; Shi, J. Hollow-structured mesoporous materials: Chemical synthesis, functionalization and applications. Adv. Mater. 2014, 26, 3176-205.

35. Purbia, R.; Paria, S. Yolk/shell nanoparticles: Classifications, synthesis, properties, and applications. Nanoscale 2015, 7, 19789-19873. 36. Liu, R.; Priestley, R. D. Rational design and fabrication of core–shell nanoparticles through a one-step/pot strategy. J. Mater. Chem. A 2016, 4, 6680-6692. 37. El-Toni, A. M.; Habila, M. A.; Labis, J. P.; ZA, A. L.; Alhoshan, M.; Elzatahry, A. A.; Zhang, F. Design, synthesis and applications of core-shell, hollow core, and nanorattle multifunctional nanostructures. Nanoscale 2016, 8, 2510-2531. 38. Lee, J.; Park, J. C.; Song, H. A nanoreactor framework of a Au@SiO2 yolk/shell structure for catalytic reduction of p-nitrophenol. Adv. Mater. 2008, 20, 1523-1528. 39. Qiao, Z. A.; Zhang, P.; Chai, S. H.; Chi, M.; Veith, G. M.; Gallego, N. C.; Kidder, M.; Dai, S. Lab-in-a-shell: Encapsulating metal clusters for size sieving catalysis. J. Am. Chem. Soc. 2014, 136, 11260-11263. 40. Liu, J.; Yang, H. Q.; Kleitz, F.; Chen, Z. G.; Yang, T.; Strounina, E.; Lu, G. Q. M.; Qiao, S. Z. Yolk-shell hybrid materials with a periodic mesoporous organosilica shell: Ideal nanoreactors for selective alcohol oxidation. Adv. Funct. Mater. 2012, 22, 591599. 41. Peng, J.; Lan, G.; Guo, M.; Wei, X.; Li, C.; Yang, Q. Fabrication of efficient hydrogenation nanoreactors by modifying the freedom of ultrasmall platinum nanoparticles within yolkshell nanospheres. Chem. Eur. J. 2015, 21, 10490-10496. 42. Subramanian, V.; Cheng, K.; Lancelot, C.; Heyte, S.; Paul, S.; Moldovan, S.; Ersen, O.; Marinova, M.; Ordomsky, V. V.; Khodakov, A. Y. Nanoreactors: An efficient tool to control the chain-length distribution in fischer–tropsch synthesis. ACS Catal. 2016, 6, 1785-1792. 43. Kuwahara, Y.; Ando, T.; Kango, H.; Yamashita, H. Palladium nanoparticles encapsulated in hollow titanosilicate spheres as an ideal nanoreactor for one-pot oxidation. Chem. Eur. J. 2017, 23, 380-389. 44. Fang, X.; Chen, C.; Liu, Z.; Liu, P.; Zheng, N. A cationic surfactant assisted selective etching strategy to hollow mesoporous silica spheres. Nanoscale 2011, 3, 1632-1639. 45. Kuwahara, Y.; Sumida, Y.; Fujiwara, K.; Yamashita, H. Facile synthesis of yolk-shell nanostructured photocatalyst with improved adsorption properties and molecular-sieving properties. ChemCatChem 2016, 8, 2781-2788. 46. Kuwahara, Y.; Furuichi, N.; Seki, H.; Yamashita, H. One-pot synthesis of molybdenum oxide nanoparticles encapsulated in hollow silica spheres: An efficient and reusable catalyst for epoxidation of olefins. J. Mater. Chem. A 2017, 5, 18518-18526. 47. Yang, H.; Zhang, L.; Zhong, L.; Yang, Q.; Li, C. Enhanced cooperative activation effect in the hydrolytic kinetic resolution of epoxides on [Co(salen)] catalysts confined in nanocages. Angew. Chem. Int. Ed. 2007, 46, 6861-6865. 48. Yang, Q.; Han, D.; Yang, H.; Li, C. Asymmetric catalysis with metal complexes in nanoreactors. Chem. Asian J. 2008, 3, 12141229. 49. Qiao, Z. A.; Huo, Q.; Chi, M.; Veith, G. M.; Binder, A. J.; Dai, S. A "ship-in-a-bottle" approach to synthesis of polymer dots@silica or polymer dots@carbon core-shell nanospheres. Adv. Mater. 2012, 24, 6017-21. 50. Yin, Y.; Chen, M.; Zhou, S.; Wu, L. A general and feasible method for the fabrication of functional nanoparticles in mesoporous silica hollow composite spheres. J. Mater. Chem. 2012, 22, 11245-11251. 51. Yue, Q.; Zhang, Y.; Wang, C.; Wang, X.; Sun, Z.; Hou, X.-F.; Zhao, D.; Deng, Y. Magnetic yolk–shell mesoporous silica microspheres with supported au nanoparticles as recyclable high-performance nanocatalysts. J. Mater. Chem. A 2015, 3, 4586-4594.

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

52. Fang, X.; Zhao, X.; Fang, W.; Chen, C.; Zheng, N. Self-templating synthesis of hollow mesoporous silica and their applications in catalysis and drug delivery. Nanoscale 2013, 5, 22052218. 53. Wu, X.; Tan, L.; Chen, D.; He, X.; Liu, H.; Meng, X.; Tang, F. Alkylaminosilane-assisted simultaneous etching and growth route to synthesise metal nanoparticles encapsulated by silica nanorattles. Chem. Eur. J. 2012, 18, 15669-15675. 54. Qi, G.; Wang, Y.; Estevez, L.; Duan, X.; Anako, N.; Park, A.-H. A.; Li, W.; Jones, C. W.; Giannelis, E. P. High efficiency nanocomposite sorbents for CO2 capture based on amine-functionalized mesoporous capsules. Energy Environ. Sci. 2011, 4, 444452. 55. Du, X.; Yao, L.; He, J. One-pot fabrication of noble-metal nanoparticles that are encapsulated in hollow silica nanospheres: Dual roles of poly(acrylic acid). Chem. Eur. J. 2012, 18, 7878-7885. 56. Stöber, W.; Fink, A. Controlled growth of monodisperse silica spheres in the micron size range. J. Colloid Interface Sci. 1968, 26, 62-69.

Page 14 of 15

57. Groen, J. C.; Peffer, L. A. A.; Pérez-Ramı́rez, J. Pore size determination in modified micro- and mesoporous materials. Pitfalls and limitations in gas adsorption data analysis. Microporous Mesoporous Mater. 2003, 60, 1-17. 58. Mojet, B. L.; Hoogenraad, M. S.; Dillen, A. J. v.; Geus, J. W.; Koningsberger, D. C. Coordination of palladium on carbon fibrils as determined by xafs spectroscopy. J. Chem. Soc., Faraday Trans. 1997, 93, 4371-4375. 59. Mei, D.; Neurock, M.; Smith, C. M. Hydrogenation of acetylene–ethylene mixtures over Pd and Pd–Ag alloys: First-principles-based kinetic monte carlo simulations. J. Catal. 2009, 268, 181-195. 60. Kobayashi, S.; Hiroishi, K.; Tokunoh, M.; Saegusa, T. Chelating properties of linear and branched poly(ethyleneimines). Macromolecules 1987, 20, 1496-1500. 61. Kuwahara, Y.; Kang, D. Y.; Copeland, J. R.; Brunelli, N. A.; Didas, S. A.; Bollini, P.; Sievers, C.; Kamegawa, T.; Yamashita, H.; Jones, C. W. Dramatic enhancement of CO2 uptake by poly(ethyleneimine) using zirconosilicate supports. J. Am. Chem. Soc. 2012, 134, 10757-10760.

ACS Paragon Plus Environment

Page 15 of 15 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Table of Contents (TOC)

ACS Paragon Plus Environment

15